0% found this document useful (0 votes)
13 views

V R R V: Longju@mit - Edu

Uploaded by

Lewis Zhou
Copyright
© © All Rights Reserved
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
13 views

V R R V: Longju@mit - Edu

Uploaded by

Lewis Zhou
Copyright
© © All Rights Reserved
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 26

Extended Quantum Anomalous Hall States in Graphene/hBN Moiré Superlattices

Zhengguang Lu1†, Tonghang Han1†, Yuxuan Yao1†, Jixiang Yang1, Junseok Seo1, Lihan Shi1, Shenyong
Ye1, Kenji Watanabe2, Takashi Taniguchi3, Long Ju1*

1
Department of Physics, Massachusetts Institute of Technology, Cambridge, MA, USA.

2
Research Center for Electronic and Optical Materials, National Institute for Materials Science, 1-1 Namiki,
Tsukuba 305-0044, Japan

3
Research Center for Materials Nanoarchitectonics, National Institute for Materials Science, 1-1 Namiki,
Tsukuba 305-0044, Japan

*Corresponding author. Email: [email protected] †These authors contributed equally to this work.

Electrons in topological flat bands can form novel topological states driven by the correlation
effects. The penta-layer rhombohedral graphene/hBN moiré superlattice has been shown to
host fractional quantum anomalous Hall effect (FQAHE) at ~400 mK1, triggering discussions
around the underlying mechanism and the role of moiré effects2–6. In particular, novel
electron crystal states with non-trivial topology have been proposed3,4,7–14. Here we report
DC electrical transport measurement in rhombohedral penta- and tetra-layer graphene/hBN
moiré superlattices at electronic temperatures down to ~40 mK. We observed two more
FQAH states in the penta-layer devices than previously reported. In a new tetra-layer device,
we observed FQAHE at filling factors v = 3/5 and 2/3 at 300 mK. With a small bias current
and the lowest temperature, we observed a new extended quantum anomalous Hall (EQAH)
state and magnetic hysteresis, where Rxy = h/e2 and vanishing Rxx span a wide range of moiré
filling factor v from 0.5 to up to 1.3. By increasing the temperature or current, FQAHE can
be recovered---suggesting the break-down of the EQAH states and a phase transition into
the fractional quantum Hall liquid15–17. Furthermore, we observed displacement field-
induced quantum phase transitions from the EQAH states to Fermi liquid, FQAH liquid and
the likely composite Fermi liquid. Our observation establishes a new topological phase of
electrons with quantized Hall resistance at zero magnetic field, and enriches the emergent
quantum phenomena in materials with topological flat bands.

Flat electronic bands with quantum geometry have been a rich platform to explore emergent
quantum phenomena driven by intertwined correlation and topology effects. In the heterostructures
of two-dimensional materials, such bands can be engineered with great flexibilities in structures
and tuned in situ through dual-gating. Consequently, a rich spectrum of quantum phases of matter
has been discovered in these synthetic quantum materials. In the example of rhombohedral
graphene-based heterostructures, gate-tunable topological flat bands stem from the highly ordered
crystalline structure and additional proximity effects due to neighboring layers18–36. It has been
demonstrated to host correlated insulators19,20,23,25,26, orbital multiferroicity24 and
superconductivity22,31. In particular, the penta-layer rhombohedral graphene/hBN moiré
superlattice has been shown to host the fractional quantum anomalous Hall effect (FQAHE) with
six fractional states1. The underlying mechanism of integer and fractional QAHE in this system,
however, is under debate2–6. This is partially due to the weak moiré potential experienced by
electrons when exhibiting FQAHE, and the lack of an isolated moiré miniband in the single-
particle picture. These apparent puzzles, the higher temperature scale of FQAHE than FQHE, and
the highly tunable band structures in this ideal material platform leave a wide un-charted territory
to explore other emergent quantum phenomena. Among a range of theoretical proposals to
describe this system, some works suggest the possibility of a quantum anomalous Hall crystal
(QAHC) that could exist under a weak moiré potential or even in the absence of moiré effects 3,4,10–
14
, which is yet to be experimentally tested or understood.

Here we report electrical transport properties of the rhombohedral graphene/hBN moiré


superlattices at electronic temperatures down to ~40 mK. In Device 1&2, which are based on
penta-layer graphene, we observed deeper dips in Rxx at fractional fillings and two more FQAH
states than previously reported1. In Device 3, which is based on tetra-layer graphene, we observed
two fractional states that realized the FQAHE. These observations reinforced the FQAHE in this
material system. At even lower temperatures and certain ranges of D, surprisingly, we observed
quantized Hall resistance Rxy = h/e2 in a wide range of moiré filling factor v from 0.5 to up to 1.3
accompanied by vanishing longitudinal resistance Rxx. These states realize the integer quantum
anomalous Hall effect1,33,37–42 in an extended range of charge density and filling factors (thus
named EQAH states), but are distinct from Chern insulators and re-entrant quantum Hall insulators
in several aspects. At increased temperature or bias current, these states are broken down and
replaced by certain FQAH states. We will discuss the nature of this newly emerged topological
state in comparison to the QAHC, re-entrant quantum Hall insulators and Wigner crystals. The
data presented in the main text is based on device 1&3, and additional data from device 2&3 are
included in Extended Data Figures. Summary of the device information, including optical image,
layer number and hBN alignment angle is shown in extended Figure 4.

FQAHE in Penta- and Tetra-layer Graphene

Figure 1a shows schematics of all three devices, in which FQAHE are observed when electrons
are polarized to the layer furthest from the moiré superlattice. We have improved the filtering of
the electronic noise and electron thermalization in our dilution refrigerator, and we believe that an
electronic temperature of ~40 mK can be reached when the mixing chamber temperature is at the
base of 10 mK.
We first examine the FQAHE in Device 1. Figure 1b shows Rxx in Device 1, the penta-
layer rhombohedral graphene/hBN moiré device, at D/𝜀! = 0.925 V/nm at 0.3, 0.2 and 0.1 K.
Clear dips can be seen at fractional filling factors v = 2/5, 3/7, 4/9, 4/7 and 3/5 as reported
previously, while new dips can be identified at v = 5/11 and 5/9 at 100 mK. At these filling factors,
the value of Rxx decreases by up to 2.5 times as the mixing chamber temperature is lowered from
0.3 K to 0.1 K.

The observations are aligned with the expectation of fractional quantum (anomalous) Hall
states, in which the Rxx decreases as the temperature is lowered and the thermal activation across
the FQ(A)H gaps is suppressed. In addition, a lower electronic temperature indeed helps revealing
more fractional states that have larger denominators and smaller energy gaps than previously
reported. Based on the temperature dependence of Rxx, we think the lowest electronic temperature
in the previous measurement on these two samples was between 0.2 and 0.3 K.

We further examine a newly made tetra-layer graphene/hBN moiré superlattice (Device 3),
the optical image of which is shown in Fig. 1c. We have optimized the device design and
fabrication so that all the six electrodes have good electrical contact to the channel at large D, and
standard Hall bar measurement is done. At a temperature of 0.3 K, quantized anomalous Hall
resistance Rxy = h/ve2 and dips in Rxx are clearly seen at v = 3/5, 2/3 and 1, as shown in Fig. 1d. The
value of Rxx at v = 2/3 is around 0.1 h/e2, even smaller than that in Device 1&2. The Rxx and Rxy
maps, and the Landau fan data of Device 3 are included as Extended Data Figure 6-8.

The data in Fig. 1d demonstrates the FQAHE in the tetra-layer graphene/hBN moiré
superlattice. Although the graphene layer number is different from the penta-layer devices, the
general phase diagram is similar to that of the latter (see Ref.1 and Extended Data Figure 6). This
observation agrees with theoretical calculations2–4,6, indicating that rhombohedral graphene/hBN
moiré superlattice is a family of materials hosting FQAHE. We note that the twist angles are
slightly different in Device 1&2 (~0.77 deg) and Device 3 (~0.55 deg), which results in the
difference between the charge densities corresponding to v = 1. This difference might explain the
difference in the number of fractional states observed in tetra- and penta-layer devices (2 vs 8). In
general, the effect of twist angle is to be examined experimentally by systematically fabricating
and measuring more devices.
Figure 1. Fractional quantum anomalous Hall effects in rhombohedral penta- and tetra-layer
graphene/hBN moiré superlattice devices. a. Schematics of the device configuration and
measurement conditions. In both penta- and tetra-layer devices, electrons are polarized to the
layer that is furthest from the moiré superlattice when the FQAHE is observed. b. Temperature-
dependent Rxx of Device 1 the penta-layer device. Clear dips can be seen at v = 2/5, 3/7, 4/9, 4/7
and 3/5 as reported previously, while new fractional states can be identified at v = 5/11, and 5/9
at the lowest temperature. At a fixed v corresponding to the FQAH state, Rxx decreases as the
temperature is lowered. c. Optical image of Device 3 the tetra-layer device. ’tg’ and ’bg’
correspond to top gate and bottom gate electrodes, respectively. Scale bar: 4 µm. We pass a
current through ’S’ & ’D’, measure Vxx between ’1’ & ’2’, and measure Vxy between ’2’ & ’3’. d.
Rxx and Rxy in the tetra-layer device at 300 mK. Quantized Rxy and dips in Rxx can be seen at v =
3/5 and 2/3---underscoring the universal FQAHE phenomena in rhombohedral graphene/hBN
moiré superlattice systems. All data are extracted by symmetrizing and anti-symmetrizing data
taken at out-of-plane magnetic fields B = ± 0.1 T.

Extended Quantum Anomalous Hall States

Figure 2 shows the phase diagram measured in Device 1 in a large range of v and gate-displacement
field D. Three regions with uniform vanishing Rxx and non-zero Rxy values can be identified in Fig.
2a&b. The first one is a diamond-shaped region at v = 1/243, which is neighbored by a big second
region covering from v = 0.55 to 0.9. The latter almost connects to the third region that covers
from v = 0.93 to 1.03. These three regions compose a big anomalous Hall region, which sandwiches
the FQAHE region together with the very insulating (and likely Wigner crystal1) region at low v
and D. Figure 2c&d reveal temperature-dependent quantitative details along the dashed lines in
Fig. 2a&b, which cut through the three aforementioned regions. At the base temperature of 10 mK
(mixing chamber temperature), Rxx almost vanishes at all filling factors from v = 1/2 to 1. At the
same time, Rxy shows a wide plateau at h/e2 covering the same range of v. There are two small gaps
between the three regions, but the temperature dependence clearly suggests their destiny towards
quantized values of Rxx = 0 and Rxy = h/e2, should the electronic temperature be further lowered.
When the temperature is increased, the values of Rxx and Rxy deviate from these quantized values
and recovered to those reported previously at higher electronic temperatures1.

In a tetra-layer rhombohedral graphene/hBN moiré superlattice device (see Extended Data


Fig. 7), we observed similar states with quantized Rxy and vanishing Rxx in a wide range of v similar
to those shown in Fig. 2. There are two differences between the tetra-layer and penta-layer devices
though: 1. In the former case, regions 2 and 3 with quantized Rxy = h/e2 are merged into one region
without a gap between them; 2. The quantized Rxy = h/e2 state extends significantly beyond v = 1
and reaches v = 1.3.

The observation of quantized Rxy = h/e2 and vanishing Rxx across a wide range of filling
factor (v = 0.5 to 1 for penta-layer graphene, v = 0.5 to 1.3 for tetra-layer graphene) clearly goes
beyond the IQAHE and FQAHE physics picture1,33,35,37–42. We name these states as extended
quantum anomalous Hall (EQAH) states. At v = 3/5 and 2/3 and certain ranges of D, FQAH states
are replaced by the EQAH states in both tetra-layer and penta-layer devices, suggesting the latter
to be the true ground state at low enough temperatures for these (v, D) combinations in the second
region. Even for the third region including v = 1, the plateau of Rxy in Fig. 2d is twice wider at 10
mK than at 380 mK and includes more states at v < 1. At continuously varying v from 0.5 to 1.3
(other than 3/5, 2/3 and 1), the quantization of Rxy and Rxx suggest that EQAH states bear a universal
feature that is not necessarily related to the underlying moiré superlattice. Even for the smallest
(region 1) among the three regions, the range of charge density and filling factor it covers is much
broader than any FQAH state covers.

Figure 2e&f show the magnetic hysteresis behaviors of Rxx and Rxy at 10 mK at two
representative positions on the phase diagram in the first and second quantized regions. Although
with very different charge densities and filling factors, both states show the quantization of Rxy =
± h/e2 at saturation. Together with data in Fig. 2a-d, we conclude that the EQAH state is a new
topological state that realizes the integer quantum anomalous Hall effect37.
a b e

D/ε0 = 0.928V/nm
B
ν = 0.719

Rxx, Rxy (h/e2 )


x 0.2

f
D/ε0 = 0.946V/nm
c d ν = 0.5
B
0.01 0.38 0.01 0.38

Rxx, Rxy (h/e2 )


T (K) T (K)
1/2 3/5 2/3 1
x 0.2

1/2 3/5 2/3 1


B

Figure 2. Extended quantum anomalous Hall states in Device 1 the penta-layer device. a&b.
Mapping of Rxx and Rxy in a large range of v and D at a mixing chamber temperature of 10 mK.
Three regions (labeled by numbers and arrows) show quantized Rxy at h/e2and vanishing Rxx,
located at around v = 1/2, spanning between v = 0.55 and 0.83, as well as around v = 1. These
three regions are almost connected into one big region with Rxy = h/e2 that swamps the FQAH
states at v > 1/2. c&d. Temperature-dependent Rxx and Rxy taken along the dashed line in a&b.
From high temperature to low temperature, Rxx decreases and the dips corresponding to FQAH
states at v = 3/5 and 2/3 become small bumps, while the Rxy approaches h/e2 from values both
below and above it. The plateau of Rxy at h/e2 spans almost from v = 1/2 to 1 at 10 mK, and the
gap at v = 0.94 and 0.53 are on the trend of being closed as the temperature is lowered. e&f.
Magnetic hysteresis scans of Rxx and Rxy at (v,𝐷 ⁄𝜀! )= (0.719, 0.928 V/nm) and (0.5, 0.946 V/nm)
respectively at 10mK, featuring quantized Rxy = h/e2 in both cases. All data are extracted by
symmetrizing/anti-symmetrizing data taken at B = ± 0.1 T.

Current-Induced Break-down

Figure 3 shows the behaviors of EQAH states under varied current excitations. We apply a DC
current through the sample and limit the AC current excitation to 50 pA to measure the
corresponding differential resistances Rxy and Rxx. Figure 3a&b show the results taken along the
dashed lines in Fig. 2a&b at 10 mK. By increasing the DC current IDC from 0 to 2.3 nA, the wide
plateaus of Rxy = h/e2 and Rxx = 0 gradually disappear and eventually evolve into shapes that are
generally similar to the curves at 380 mK in Fig. 2c&d and in our previous work1.

Phenomenologically, increasing the DC current has the same effect of weakening the
EQAH states as increasing the temperature does. At IDC = 2.3 nA, both Rxy and Rxx curves mimic
in shape those in Fig. 2c&d at high temperatures, and the value of Rxy recovers to that of the
FQAHE at v = 3/5 and 2/3. These observations suggest that the EQAH ground states is broken-
down and replaced by the fractional quantum Hall liquids15–17 under high DC current excitations.

However, there are also important differences between the Rxy and Rxx curves at elevated
temperature and increased DC current. In Fig. 2c, Rxx increases almost monotonically as the
temperature is increased at all filling factors. In Fig. 3b, however, Rxx first increases then decreases
as the DC current increases. This non-monotonic trend is most obvious at v = 0.6-0.8 and can also
be seen for other filling factors. Such non-monotonic dependence on the DC current can also be
seen for Rxy in Fig. 3a. We highlight such non-monotonic behaviors of Rxy and Rxx at two
representative states that show Rxy = h/e2 and Rxx = 0, as can be seen in Fig. 3c-f. The first state
resides in the second region of quantized Rxy while the second state resides inside the diamond-
shaped first region of quantized Rxy. In both cases, Rxx exhibits clear threshold behaviors: being
zero at small current and showing a pair of peaks at a critical value of current. Correspondingly,
Rxy exhibits the quantized value of h/e2 at small current and sudden changes at the same critical
current as in the Rxx curve. Such variations of Rxy and Rxx as a function of the DC current disappear
at high temperatures, where both Rxy and Rxx remain constant.

Such threshold and non-monotonic behaviors of Rxy and Rxx indicate non-linear voltage-
current relations. The observations are reminiscent of two other well-known examples of
correlated electron ground states. Firstly, Fig. 3d resembles the differential resistance behavior in
an s-wave superconductor: the zero-resistance Cooper pairs carry the current until hitting a critical
value and a pair of resistance peaks emerge. In the case of EQAH states, the current is carried by
the chiral edge states until reaching a critical value and the bulk transport starts to contribute.
Secondly, such break-down behaviors suggested by the differential resistances are also reminiscent
of the Chern insulators that exhibit integer quantum anomalous Hall effects44–47. In Chern
insulators with the QAHE, the insulating bulk of the sample breaks down at high current/voltage,
and transitions into a Fermi liquid state. Quantitatively, however, the breakdown of the Chern
insulator and the EQAH states are quite different. In Extended Data Fig. 8, we compare the Rxx vs
IDC for the EQAH states and the Chern insulator at v = 1 in the same tetra-layer device. The
threshold current in the EQAH state is 50 times smaller than that in the Chern insulator state of the
same device---suggesting a different mechanism of break-down from previously reported for
Chern insulators44,45.
a c e
0 2.3 0.01 0.38 0.01 0.38
T (K) T (K)
IDC (nA)

3/5 2/3
D/ε0 = 0.948V/nm D/ε0 = 0.951V/nm
1/2
ν = 0.79 ν = 0.5

b d f
0 2.3 D/ε0 = 0.948V/nm 0.01 0.38 D/ε0 = 0.951V/nm 0.01 0.38
ν = 0.79 ν = 0.5
IDC (nA) T (K) T (K)
3/5 2/3
1/2

Figure 3. Break-down of EQAH states in Device 1 the penta-layer graphene/hBN device. a&b.
Differential resistance Rxx and Rxy taken at the dashed lines in Fig. 2a&b at 10 mK, using an AC
current of 50 pA on top of a varied DC current. The quantized Rxy = h/e2 and vanishing Rxx can be
seen at small DC current, while the high-temperature behavior (including FQAHE at v = 3/5 and
2/3) as shown in Fig. 2 is qualitatively reproduced at IDC = 2.3 nA. This suggests that the EQAH
state is the true ground state at this displacement field in certain range of D, while the fractional
quantum Hall liquid can be recovered by a large DC current. Non-monotonic dependence of Rxx
on IDC, however, can be seen at intermediated IDC. This is clearly different from the monotonic
dependence of Rxx on the temperature, as shown in Fig. 2c. c&d. Current-dependent Rxx and Rxy
taken at (v, D/ 𝜀! ) = (0.74, 0.948 V/nm) and varying temperatures, revealing the quantized
transport at small current in the EQAH state, and the anomalous quantum Hall liquid transport
at high current when the EQAH state is broken down. The pair of peaks in Rxx at the break-down
threshold corresponds to the non-monotonic dependence of Rxx on IDC in b, and is reminiscent of
s-wave superconductors and Chern insulators to some extent. e&f. Current-dependent Rxx and Rxy
taken at (v, D/𝜀! ) = (0.5, 0.951 V/nm) and varying temperatures, showing similar behaviors as in
c&d.

Displacement Field-Induced Phase Transitions

The displacement field D provides an important tuning knob of the flat band physics in
rhombohedral graphene. With a small change of D, one can fine-tune the band structure to
influence the competition between different ground states that are close in energy. We now
examine the effect of D on the observed EQAH states, as well as other electron liquid states in the
same device.
Figure 4a&b show the Rxy and Rxx as a function of D. At around D/𝜀! = 0.92 V/nm, Rxy
shows a value of 2h/e2, which is compatible with the composite Fermi liquid (CFL) picture. At
high temperatures, Rxy decreases from 2h/e2 gradually while Rxx non-monotonically as D/ 𝜀!
changes from 0.9 to 0.94 V/nm. At low temperatures, a plateau at Rxy = h/e2 and Rxx = 0 emerges
at intermediate Ds. This observation is consistent with the EQAH state in the diamond-shaped
region at low temperatures, and indicates quantum phase transitions from the CFL to the EQAH
state to the valley-polarized Fermi liquid (FL) triggered by D.

Figure 4c&d show Rxy and Rxx as a function of D at varying DC current bias. Again, the
plateaus at Rxy = h/e2 and Rxx = 0 emerge at intermediate D only when the IDC is small. Such
dependence of the differential resistances on current does not show up for the CFL, as can be seen
in Fig. 4e&f, where both Rxy and Rxx are constants regardless of the temperature. In contrast, in the
range of D corresponding to the EQAH state, the breakdown happens only at low temperatures but
not high temperatures, as shown in Fig. 4g&h. The differences in response to temperature and
current clearly mark the different nature of the likely-to-be CFL and the EQAH state.

At some fractional fillings where FQAHE were observed, D can also induce quantum phase
transitions. This is shown in in Fig. 4i-k, in which the filling factor is fixed at v = 3/5, 4/7, 5/9,
respectively48. At the base temperature of 10 mK, FQAH state and EQAH state both occupy a
range of D, as evidenced by the plateaus at Rxy = h/ve2 and Rxy = h/e2. Further tuning of D can
induce phase transitions to Fermi liquid and Wigner crystals at the high and low ends of D,
respectively.
FQAH
a WC CFL NQAH FL
e i WC NQAH FL

D/ε0 = 0.921V/nm 5/3


ν = 0.5

Rxx, Rxy (h/e2)


1

0.01 0.38
T (K)
T (K)
0.01 0.38
f
b D/ε0 = 0.921V/nm
0.01 0.38 ν = 0.5 0.90 0.92 0.94 0.96
T (K) D/ε0 (V/nm)
j
0.01 0.38
7/4
T (K)

Rxx, Rxy (h/e2)


0.88 0.90 0.92 0.94 0.96
1
D/ε0 (V/nm)
c IDC (nA)
0 2.3 g
D/ε0 = 0.946V/nm
ν = 0.5

0.90 0.92 0.94 0.96


D/ε0 (V/nm)
0.01 0.38 k
T (K) 9/5

h Rxx, Rxy (h/e2)


D/ε0 = 0.946V/nm
d ν = 0.5 1
0.01 0.38
T (K)

0.90 0.92 0.94 0.96


0.88 0.90 0.92 0.94 0.96
D/ε0 (V/nm)
D/ε0 (V/nm)

Figure 4. Phase transitions from the EQAH states to (composite) Fermi liquid and FQAH
liquid. a&b. Rxy and Rxx at v = 1/2 and varying D in a range of temperatures. The plateaus at Rxy
= 2h/e2 and h/e2 that correspond to CFL and EQAH can be clearly seen at low temperatures. At
very high D, the transport behavior agrees with a valley-polarized Fermi-liquid. At very low D,
the resistance diverges and suggests a Wigner crystal state. c&d. Differential resistances Rxy and
Rxx taken at v = 1/2 and varying D at 10 mK, using an AC current of 50 pA on top of a DC current.
Behaviors in different ranges of D qualitatively agree with that in a&b. e&f. Current dependence
of differential resistances Rxy and Rxx at (v, D/𝜀! ) = (0.5, 0.921 V/nm), agreeing with the linear
voltage-current transport of CFL. g&h. Current dependence of differential resistances Rxy and Rxx
at (v, D/𝜀! ) = (0.5, 0.946 V/nm), revealing the linear voltage-current transport behavior at high
temperatures when the EQAH state is broken-down to fractional quantum Hall liquid, which is
contrasted by the non-linear voltage-current transport when the EQAH is un-perturbed at low
temperatures. i-k. Dependence of Rxy and Rxx on D at 10 mK, taken at v = 3/5, 4/7 and 5/9
respectively. Plateaus corresponding to FQAH and EQAH states are clearly seen and connected
by phase transitions induced by D at the same v.
Discussion

Our observation of quantized Rxy and vanishing Rxx in a wide range of v = 0.5 to up to 1.3 at zero
magnetic field, the temperature dependences of resistances, and the strong threshold behavior in
transport at small current indicate that the EQAH state is a new topological phase of correlated
electrons. Here we discuss more on the nature of this EQAH state with two possible underlying
pictures.

In picture 1, what is observed could be similar to the quantum anomalous Hall crystal
(QAHC) suggested by several recent theory works on the rhombohedral graphene3,4,11,12. QAHC
could exist in the absence of moiré effect and would break the time-reversal-symmetry and
translational symmetry spontaneously and simultaneously. The experimentally observed EQAH
prevails in a wide range of electron density and moiré filling factor, and most of them are in-
commensurate with the moiré superlattice. This is distinct from the generalized Wigner crystal
which only shows up at fractional filling factors of the moiré superlattice49–52. Therefore, the role
of moiré in the formation of the EQAH state is likely different from that in the formation of
generalized Wigner crystals (or charge density wave). It is possible that the weak but non-zero
moiré potential (through a remote Coulomb potential, for example) could facilitate the formation
of QAHC by modulating the band structure and Berry curvature distribution. Experimentally, an
ideal experiment to test the QAHC picture is to measure rhombohedral graphene devices in the
absence of moiré effect (for example, at large twist angles between graphene and hBN). We note
that in order to stabilize the QAHC in the absence of moiré, it is likely that even lower electronic
temperatures, smaller bias current, and the right amount of impurities (to pin the electron crystal
as in the case of Wigner crystals) will be needed. These conditions are to be carefully engineered
in future experiments.

In picture 2, the observed EQAH state could be similar to the re-entrant quantum Hall effect
(RQHE) state in two-dimensional electron gas at high magnetic fields53–59. The RQHE can be
generally understood as the superposition of a gapped integer quantum Hall liquid17 at integer
filling of Landau levels and a Wigner crystal formed by the excess charges. In analogy, topological
flat bands play the role of Landau levels and the integer QAHE can be contributed by a QAH liquid
at v = 1. The charges corresponding to (1-v) will have to form a Wigner crystal that contributes a
diverging Rxx and zero Rxy. In this case, although it is not entirely clear how moiré superlattice
drive the QAHE at v = 1, some kind of moiré effect is needed for the formation of EQAH states as
we observed. However, we note that the RQHE happens predominantly in high-index Landau
levels (with rare exceptions 54,59), while the EQAH state we observed happens mostly in the lowest
moiré band. RQHE was also never observed for a large and continuous filling range as we observed
for the EQAH states. Especially, RQHE never kills fractional quantum Hall states, while EQAH
states replace some FQAH states at the base temperature.
In either case, our observed EQAH state is a new topological phase of electrons that is
distinct from other topological states so far: 1. It happens spontaneously at zero magnetic field and
is thus different from all states exhibiting quantum Hall effects under a magnetic field; 2. It
happens in a wide range of moiré filling factor and is thus different from all Chern insulators that
are tied to specific integer or fraction fillings. Both the two pictures above and the threshold
behavior of electric transport suggest an underlying electron crystal in the EQAH state. In two-
dimensional electron gases, topologically trivial electron crystals have been realized when the
effective mass is large60–63, by quenching the kinetic energy under high magnetic fields64–69, or be
stabilized by the commensuration with an underlying moiré superlattice or another layer hosting
electrons49–51. A topological electron crystal at zero magnetic field, however, has never been
observed in any experiment. Further experiments are needed to test the possible electron crystal
nature of our observed EQAH state, through direct imaging in the real space50,62,69, measuring
electron diffraction peaks in the momentum space70, or by measuring narrow-band noise71–73.
These measurements would require more advanced device engineering, better control of electronic
temperature, and novel microscopy experiments, which are beyond the scope of this work. In the
meanwhile, our experiment opens up opportunities to study the QAHC, an unprecedented
topological state of matter that breaks both the time-reversal and translational symmetries
spontaneously.

References

1. Lu, Z. et al. Fractional quantum anomalous Hall effect in multilayer graphene. Nature 2024
626:8000 626, 759–764 (2024).

2. Dong, Z., Patri, A. S. & Senthil, T. Theory of fractional quantum anomalous Hall phases in
pentalayer rhombohedral graphene moir\’e structures. arXiv: 2311.03445 (2023).

3. Zhou, B., Yang, H. & Zhang, Y.-H. Fractional quantum anomalous Hall effects in rhombohedral
multilayer graphene in the moir\’eless limit and in Coulomb imprinted superlattice. arXiv:
2311.04217 (2023).

4. Dong, J. et al. Anomalous Hall Crystals in Rhombohedral Multilayer Graphene I: Interaction-


Driven Chern Bands and Fractional Quantum Hall States at Zero Magnetic Field. arXiv:
2311.05568 (2023).

5. Guo, Z., Lu, X., Xie, B. & Liu, J. Theory of fractional Chern insulator states in pentalayer graphene
moir\’e superlattice. arXiv: 2311.14368 (2023).

6. Kwan, Y. H. et al. Moir\’e Fractional Chern Insulators III: Hartree-Fock Phase Diagram, Magic
Angle Regime for Chern Insulator States, the Role of the Moir\’e Potential and Goldstone Gaps in
Rhombohedral Graphene Superlattices. arXiv: 2312.11617 (2023).

7. Halperin, B. I., Teanović, Z. & Axel, F. Compatibility of Crystalline Order and the Quantized Hall
Effect. Phys Rev Lett 57, 922 (1986).
8. Teanović, Z., Axel, F. & Halperin, B. I. ‘“Hall crystal”’ versus Wigner crystal. Phys Rev B 39, 8525
(1989).

9. Kivelson, S., Kallin, C., Arovas, D. P. & Schrieffer, J. R. Cooperative ring exchange and the
fractional quantum Hall effect. Phys Rev B 36, 1620 (1987).

10. Zeng, Y., Guerci, D., Crépel, V., Millis, A. J. & Cano, J. Sublattice structure and topology in
spontaneously crystallized electronic states. Phys Rev Lett 132, 236601 (2024).

11. Soejima, T. et al. Anomalous Hall Crystals in Rhombohedral Multilayer Graphene II: General
Mechanism and a Minimal Model. arXiv: 2403.05522 (2024).

12. Dong, Z., Patri, A. S. & Senthil, T. Stability of Anomalous Hall Crystals in multilayer rhombohedral
graphene. arXiv: 2403.07873 (2024).

13. Tan, T. & Devakul, T. Parent Berry curvature and the ideal anomalous Hall crystal. arXiv:
2403.04196 (2024).

14. Sheng, D. N., Reddy, A. P., Abouelkomsan, A., Bergholtz, E. J. & Fu, L. Quantum anomalous Hall
crystal at fractional filling of moir\’e superlattices. Phys Rev Lett 133, 066601 (2024).

15. Laughlin, R. B. Anomalous quantum Hall effect: An incompressible quantum fluid with fractionally
charged excitations. Phys Rev Lett 50, 1395–1398 (1983).

16. Eisenstein, J. P. & Stormer, H. L. The Fractional Quantum Hall Effect. Science (1979) 248, 1510–
1516 (1990).

17. Kivelson, S., Lee, D. H. & Zhang, S. C. Global phase diagram in the quantum Hall effect. Phys Rev B
46, 2223 (1992).

18. Bao, W. et al. Stacking-dependent band gap and quantum transport in trilayer graphene. Nat.
Phys. 7, 948–952 (2011).

19. Myhro, K. et al. Large tunable intrinsic gap in rhombohedral-stacked tetralayer graphene at half
filling. 2d Mater 5, 045013 (2018).

20. Shi, Y. et al. Electronic phase separation in multilayer rhombohedral graphite. Nature 2020
584:7820 584, 210–214 (2020).

21. Zhou, H. et al. Half- and quarter-metals in rhombohedral trilayer graphene. Nature 2021
598:7881 598, 429–433 (2021).

22. Zhou, H., Xie, T., Taniguchi, T., Watanabe, K. & Young, A. F. Superconductivity in rhombohedral
trilayer graphene. Nature 2021 598:7881 598, 434–438 (2021).

23. Han, T. et al. Correlated Insulator and Chern Insulators in Pentalayer Rhombohedral Stacked
Graphene. Nat. Nanotechnology 19, 181–187 (2023).

24. Han, T. et al. Orbital Multiferroicity in Pentalayer Rhombohedral Graphene. Nature 623, 41–47
(2023).
25. Liu, K. et al. Interaction-driven spontaneous broken-symmetry insulator and metals in ABCA
tetralayer graphene. Nat Nanotechnol 19, 188–195 (2023).

26. Lee, Y. et al. Competition between spontaneous symmetry breaking and single-particle gaps in
trilayer graphene. Nat. Commun. 5, (2014).

27. Zou, K., Zhang, F., Clapp, C., MacDonald, A. H. & Zhu, J. Transport studies of dual-gated ABC and
ABA trilayer graphene: Band gap opening and band structure tuning in very large perpendicular
electric fields. Nano Lett 13, 369–373 (2013).

28. Zhang, L., Zhang, Y., Camacho, J., Khodas, M. & Zaliznyak, I. The experimental observation of
quantum Hall effect of l = 3 chiral quasiparticles in trilayer graphene. Nat. Phys. 7, 953–957
(2011).

29. Chen, G. et al. Evidence of a gate-tunable Mott insulator in a trilayer graphene moiré
superlattice. Nature Physics 2019 15:3 15, 237–241 (2019).

30. Chen, G. et al. Tunable correlated Chern insulator and ferromagnetism in a moiré superlattice.
Nature 579, 56–61 (2020).

31. Chen, G. et al. Signatures of tunable superconductivity in a trilayer graphene moiré superlattice.
Nature 572, 215–219 (2019).

32. Sha, Y. et al. Observation of a Chern insulator in crystalline ABCA-tetralayer graphene with spin-
orbit coupling. Science (1979) 384, 414–419 (2024).

33. Han, T. et al. Large quantum anomalous Hall effect in spin-orbit proximitized rhombohedral
graphene. Science 384, 647–651 (2024).

34. Zhou, W. et al. Layer-polarized ferromagnetism in rhombohedral multilayer graphene. Nature


Communications 2024 15:1 15, 1–8 (2024).

35. Xie, J. et al. Even- and Odd-denominator Fractional Quantum Anomalous Hall Effect in Graphene
Moire Superlattices. arXiv: 2405.16944 (2024).

36. Ding, J. et al. Electrical switching of chirality in rhombohedral graphene Chern insulators. arXiv:
2406.14289 (2024).

37. Chang, C. Z. et al. Experimental observation of the quantum anomalous Hall effect in a magnetic
topological insulator. Science (1979) 340, 167–170 (2013).

38. Serlin, M. et al. Intrinsic quantized anomalous Hall effect in a moiré heterostructure. Science
(1979) 367, 900–903 (2020).

39. Li, T. et al. Quantum anomalous Hall effect from intertwined moiré bands. Nature 600, 641–646
(2021).

40. Park, H. et al. Observation of Fractionally Quantized Anomalous Hall Effect. Nature 622, 74–79
(2023).
41. Xu, F. et al. Observation of Integer and Fractional Quantum Anomalous Hall Effects in Twisted
Bilayer MoTe2. Phys Rev X 13, 031037 (2023).

42. Su, R. et al. Generalized anomalous Hall crystals in twisted bilayer-trilayer graphene. arXiv:
2406.17766 (2024).

43. Sheng, D. N., Reddy, A. P., Abouelkomsan, A., Bergholtz, E. J. & Fu, L. Quantum anomalous Hall
crystal at fractional filling of moir\’e superlattices. arXiv: 2402.17832 (2024).

44. Kawamura, M. et al. Current-Driven Instability of the Quantum Anomalous Hall Effect in
Ferromagnetic Topological Insulators. Phys Rev Lett 119, 016803 (2017).

45. Fox, E. J. et al. Part-per-million quantization and current-induced breakdown of the quantum
anomalous Hall effect. Phys Rev B 98, 075145 (2018).

46. Zeng, Z. et al. Two-dimensional magneto-quantum transport on GaAs-AlxGa1-xAs


heterostructures under non-ohmic conditions. Journal of Physics C: Solid State Physics 16, 5441
(1983).

47. Komiyama, S. & Kawaguchi, Y. Heat instability of quantum Hall conductors. Phys Rev B 61, 2014
(2000).

48. Song, X. Y., Jian, C. M., Fu, L. & Xu, C. Intertwined fractional quantum anomalous Hall states and
charge density waves. Phys Rev B 109, 115116 (2024).

49. Regan, E. C. et al. Mott and generalized Wigner crystal states in WSe2/WS2 moiré superlattices.
Nature 2020 579:7799 579, 359–363 (2020).

50. Li, H. et al. Imaging two-dimensional generalized Wigner crystals. Nature 597, 650–654 (2021).

51. Xu, Y. et al. Correlated insulating states at fractional fillings of moiré superlattices. Nature 2020
587:7833 587, 214–218 (2020).

52. Polshyn, H. et al. Topological charge density waves at half-integer filling of a moiré superlattice.
Nature Physics 2021 18:1 18, 42–47 (2021).

53. Xia, J. S. et al. Electron correlation in the second landau level: A competition between many
nearly degenerate quantum phases. Phys Rev Lett 93, 176809 (2004).

54. Liu, Y. et al. Observation of reentrant integer quantum Hall states in the lowest Landau level.
Phys. Rev. Lett. 109, 036801 (2012).

55. Eisenstein, J. P., Cooper, K. B., Pfeiffer, L. N. & West, K. W. Insulating and Fractional Quantum Hall
States in the First Excited Landau Level. Phys Rev Lett 88, 076801 (2002).

56. Kumar, A., Csáthy, G. A., Manfra, M. J., Pfeiffer, L. N. & West, K. W. Nonconventional odd-
denominator fractional quantum hall states in the second landau level. Phys Rev Lett 105,
246808 (2010).

57. Deng, N. et al. Collective nature of the reentrant integer quantum Hall states in the second
Landau level. Phys Rev Lett 108, 086803 (2012).
58. Chen, S. et al. Competing fractional quantum Hall and electron solid phases in graphene. Phys.
Rev. Lett. 122, 026802 (2019).

59. Zhou, H., Polshyn, H., Taniguchi, T., Watanabe, K. & Young, A. F. Solids of quantum Hall skyrmions
in graphene. Nature Physics 2019 16:2 16, 154–158 (2019).

60. Smoleński, T. et al. Signatures of Wigner crystal of electrons in a monolayer semiconductor.


Nature 2021 595:7865 595, 53–57 (2021).

61. Zhou, Y. et al. Bilayer Wigner crystals in a transition metal dichalcogenide heterostructure.
Nature 2021 595:7865 595, 48–52 (2021).

62. Xiang, Z. et al. Quantum Melting of a Disordered Wigner Solid. arXiv: 2402.05456 (2024).

63. Falson, J. et al. Competing correlated states around the zero-field Wigner crystallization
transition of electrons in two dimensions. Nature Materials 2022 21:3 21, 311–316 (2021).

64. Grimes, C. C. & Adams, G. Evidence for a liquid-to-crystal phase transition in a classical, two-
dimensional sheet of electrons. Phys. Rev. Lett. 42, 795–798 (1979).

65. Andrei, E. Y. et al. Observation of a magnetically induced Wigner solid. Phys. Rev. Lett. 60, 2765–
2768 (1988).

66. Goldman, V. J., Santos, M., Shayegan, M. & Cunningham, J. E. Evidence for two-dimensional
quantum Wigner crystal. Phys. Rev. Lett. 65, 2189–2192 (1990).

67. Jiang, H. W. et al. Quantum liquid versus electron solid around ν = 1/5 Landau-level filling. Phys.
Rev. Lett. 65, 633–636 (1990).

68. Santos, M. B. et al. Observation of a reentrant insulating phase near the 1/3 fractional quantum
Hall liquid in a two-dimensional hole system. Phys. Rev. Lett. 68, 1188–1191 (1991).

69. Tsui, Y. C. et al. Direct observation of a magnetic-field-induced Wigner crystal. Nature 2024
628:8007 628, 287–292 (2024).

70. Kakiuchi, T., Wakabayashi, Y., Sawa, H., Itou, T. & Kanoda, K. Wigner crystallization in (DI-
DCNQI)2Ag detected by synchrotron radiation X-ray diffraction. Phys Rev Lett 98, 066402 (2007).

71. Cooper, K. B., Eisenstein, J. P., Pfeiffer, L. N. & West, K. W. Observation of Narrow-Band Noise
Accompanying the Breakdown of Insulating States in High Landau Levels. Phys Rev Lett 90, 4
(2003).

72. Chen, Y. P. et al. Evidence for two different solid phases of two-dimensional electrons in high
magnetic fields. Phys Rev Lett 93, 206805 (2004).

73. Madathil, P. T. et al. Moving Crystal Phases of a Quantum Wigner Solid in an Ultra-High-Quality
2D Electron System. Phys Rev Lett 131, 236501 (2023).

74. Lui, C. H. et al. Imaging stacking order in few-layer graphene. Nano Lett 11, 164–169 (2011).

75. Yan, J. A., Ruan, W. Y. & Chou, M. Y. Enhanced optical conductivity induced by surface states in
ABC-stacked few-layer graphene. Phys Rev B Condens Matter Mater Phys 83, 245418 (2011).
76. Ju, L. et al. Topological valley transport at bilayer graphene domain walls. Nature 520, 650–655
(2015).

Method

High Through-Put Fabrication of Rhombohedral Graphene Stacks Enabled by Advanced


Infrared Imaging and Device Fabrication

Device 1&2 were used in Ref. 1 where the description of device fabrication can be found. Device
3 was made in generally the same procedures as Device 1&2, except for the imaging of the
rhombohedral stacking order. We developed a new infrared imaging technique based on InGaAs
camera, which is installed on a regular optical microscope and can take pictures of the graphene
flakes. Different stacking orders in graphene can be identified based on their different infrared
conductivities and contrast with the substrate, as described in Ref74,75. Compared to the near field
infrared nanoscopy technique we employed previously, the far-field imaging based on the InGaAs
camera has a much higher efficiency due to the multi-pixel data collection simultaneously. We
combine the near-field and far-field approaches, together with Raman spectroscopy to visualize
and confirm stacking orders in the exfoliated flakes and assembled stacks.

Extended Data Figure 1 shows how this imaging method is implemented. We picked up
the top hBN, graphite, middle hBN, and the penta-layer graphene using polypropylene carbonate
(PPC) film and landed it on a prepared bottom stack consisting of an hBN and graphite bottom
gate. We used the IR camera to quickly screen the exfoliated graphene flakes and identify the
rhombohedral domains. We also checked if the stacking order is preserved after picking up the
graphene with hBN and after the stack is finished. The device was then etched into a Hall bar
structure using standard e-beam lithography (EBL) and reactive-ion etching (RIE). We deposited
Cr/Au for electrical connections to the source, drain and gate electrodes.
There are two caveats to note though, as shown in Extended Data Figure 2. Firstly, for
multilayer graphene thicker than three layers, there are more intermediate stacking orders other
than rhombohedral and Bernal stacking. Some of the intermediate stackings show similar
responses to the rhombohedral stacking on the InGaAs camera and could mislead the judgment.
Near-field infrared nanoscopy works better in differentiating rhombohedral stacking from other
stackings, possibly due to the longer wavelength used in the near-field measurements. Secondly,
the spatial resolution of the far-field imagining is limited to ~one micron due to the diffraction
limit of infrared light, and could miss domain structures with mixed stacking orders and domain
walls that are smaller than the spatial resolution32,76.

Transport measurement

The device was measured in a Bluefors LD250 dilution refrigerator with an electronic temperature
of around 40 mK. Electronic temperature is estimated by a superconducting state in other graphene
device with Tc around 40mK as shown in extended Figure 3. Stanford Research Systems SR830
lock-in amplifiers were used to measure the longitudinal and Hall resistance Rxx and Rxy with an
AC frequency at 17.77 Hz. The DC and AC currents are generated by Keysight 33210A function
generator through a 300MOhm resistor. Keithley 2400 source-meters were used to apply top and
bottom gate voltages. Top-gate voltage Vt and bottom-gate voltage Vb are swept to adjust doping
density ne = (CtVt + CbVb)/e and displacement field D/e0 = (CtVt - CbVb)/2, where Ct and Cb are top-
gate and bottom-gate capacitance per area calculated from the Landau fan diagram.

Dis-entangling longitudinal and Hall resistance

Devices 1&2 are the same as used in Ref.1 and has been measured in the same way to dis-entangle
the Rxx and Rxy. As shown in Fig. 1c, device 3 is in a Hall bar geometry with good electric contacts
on both sides of the channel at high displacement fields. We use electrodes ‘S’ & ‘D’ to pass the
current, electrodes ‘1’ & ‘2’ to measure Rxx and electrodes ‘2’ & ‘3’ to measure Rxy. Measurements
performed at opposite magnetic fields (larger than the coercive field) can thus be used to extract
Rxx and Rxy at B = 0 for the IQAH, FQAH and EQAH states, following:

Rxx(0) = (R(B) + R(-B))/2 and Rxy(0) = (R(B) - R(-B))/2.

EQAH States in Device 2


Extended Data Figure 5 shows data from Device 2, the other penta-layer graphene/hBN device on
the same chip as Device 1. The twist angle and general phase diagram of Device 2 are similar to
those of Device 1, featuring FQAHE, EQAH states and phase transitions similar to described in
Fig. 1-4 in the main text.

FQAHE and EQAH States in Device 3

Extended Data Figure 6-8 show data from Device 3, a newly made tetra-layer rhombohedral
graphene/hBN device. At 300 mK, the phase diagram shows IQAHE and FQAHE. At 10 mK, the
EQAH states emerge and dominate a large range of (v, D) from v = 0.5 to 1.3.

Author Contributions

L.J. supervised the project. Z.L. and T.H. performed the DC magneto-transport measurement.
T.H., Y.Y., L.S., and S.Y. fabricated the devices. J.Y., J.S., Z.L. and T.H. helped with installing
and testing the dilution refrigerator. K.W. and T.T. grew hBN crystals. All authors discussed the
results and wrote the paper.

Competing Interests The authors declare no competing interests.

Data availability The data that support the findings of this study are available from the
corresponding authors upon reasonable request.

Extended Data Figures


Extended Data Figure 1. Visible and infrared far-field imaging of multilayer graphene during
fabrication. a. Optical image of an exfoliated graphene flake on SiO2/Si substrate, taken with a
regular optical microscope. The majority of the flake (outlined by dashed lines) is tetra-layer.
Scale bar: 20 µm. No contrast can be seen within the tetra-layer region. b. Infrared image of the
same flake as in a, taken by an InGaAs camera. Domains with different contrasts can be seen,
which correspond to different stacking orders. The rhombohedrally stacked domain (ABCA) has
a contrast between the Bernal stacked domain (ABAB) and an intermediate stacked domain
(ABCB). These stacking orders have been confirmed by Raman scattering measurement. We cut 6
rectangles out of the ABCA domain using a laser, as outlined by dashed boxes. c. The InGaAs
camera image of 3 (out of 6 ABCA) flakes that are picked up. The two flakes on the left have
partially changed to other stacking orders, while the right-most flake remains in ABCA stacking.
d. The InGaAs camera image of the three flakes in c after they are dropped down to hBN and
bottom graphite gate. The right-most flake in c remained in the ABCA stacking, as indicated by
the right arrow. It has a clear contrast with the middle flake in c. The ABCA flake in the dashed
box was made into Device 3.

Extended Data Figure 2. Comparison between far-field infrared imaging and near-field
infrared imaging. a&b. InGaAs camera image and near-field infrared nanoscopy image of an
exfoliated multilayer graphene flake on SiO2/Si substrate. The latter reveals more domains
(labeled as 1-4) with clear contrast than the former does. c&d. Near-field infrared nanoscopy
images of two exfoliated trilayer graphene flakes. Both images show domain and domain walls
that have dimensions much smaller than 1 µm, which is well-below the diffraction-limit of the far-
field imaging based on the InGaAs camera.
Extended Data Figure 3. Estimation of the electron temperature. a. Temperature dependence of
Rxx in a graphene superconducting state. The density range of the superconducting dome remains
expanding below 40mK. b. Temperature dependence of Rxx at ne = 1.9×1012 cm-2.
Extended Data Figure 4. Summary of the device and contacts information. a. Basic information
of the three devices. b. Schematics of the gates and contacts layout in device 1 and 2. Top and
bottom graphite gate shifted relative to each other creating a n-p-n junction on one side of the
contacts. c. Gates and contacts layout in device 3 with optimized geometry. No p-n junction will
be formed on either side of the device.

a b

2/3 2/3
2/5 1/2 3/5 2/5 1/2 3/5
1.0 1.0

0.9 0.9

0.8 0.8

0.2 0.2

c d
3h/2e2

2/3
0 3 0 3
IDC (nA) IDC (nA)

0.88 0.90 0.92 0.94 0.96


D/ε0 (V/nm)

Extended Data Figure 5. Phase diagram of Device 2, a penta-layer rhombohedral


graphene/hBN moiré superlattice, at IDC = 0 A and IAC = 0.2 nA. a&b. Mapping of Rxx and Rxy
in a large range of v and D at a mixing chamber temperature of 10 mK. The main features are
similar to that of Device 1 shown in Fig. 2. Three regions show quantized Rxy at h/e2and vanishing
Rxx, located at around v = 1/2, spanning between v = 0.55 and 0.83, as well as around v = 1. These
three regions are almost connected into one big region with Rxy = h/e2 that swamps the FQAH
states at v > 1/2. c. Line-cut of Rxy at D/𝜀! = 0.96 V/nm, featuring a wide plateau at h/e2. d. Rxy at
v = ½, featuring the phase transition from CFL to EQAH driven by displacement field.

Extended Data Figure 6. FQAHE in Device 3 at 300 mK. a&b. Mapping of Rxx and Rxy in a large
range of v and D at IDC = 0 nA and IAC = 0.1 nA, which is beyond the break-down threshold current
of EQAH states. c. Landau fan corresponding to the line-cut in c, where the dashed lines are
derived from the Streda’s formula for states with Rxy = 5h/3e2, 3h/2e2 and h/e2, respectively. The
dispersions of dips in Rxx agree well with the dashed lines, as expected for fractional and integer
Chern insulators at v = 3/5, 2/3 and 1.
Extended Data Figure 7. Phase diagram of Device 3, a tetra-layer rhombohedral graphene/hBN
moiré superlattice, at 10 mK, IDC = 0 A and IAC = 0.2 nA. a&b. Mapping of Rxx and Rxy in a large
range of v and D at a mixing chamber temperature of 10 mK. Different from Device 1&2, the
region with quantized Rxy at h/e2and vanishing Rxx shifts to higher moiré filling factors. This is
likely due to the smaller twist angle and smaller charge density corresponding to v = 1 in Device
3. As a result, region 2 and 3 in Fig. 2 merge into one region without a gap in between. At the
same time, the EQAH region extends significantly beyond v = 1 and reaches v = 1.3.
Extended Data Figure 8. Current-induced break-down of EQAH state and Chern insulators in
Device 3 at 10 mK. a. Mapping of Rxy in a large range of v and D at IDC = 0 A and IAC = 0.2 nA.
Five states at ‘stars’ positions show quantized Rxy at h/e2and vanishing Rxx. b&c. Rxx and Rxy as a
function of IDC at the light and dark blue ‘star’ positions corresponding to v=1 in a, showing the
break-down behavior of the C=1 Chern insulator at large current (>50 nA). d-f. Rxx and Rxy as a
function of IDC at the green, pink and purple ‘star’ positions in a, showing the break-down behavior
of the EQAH states at small current (< 1 nA) in contrast to the large break-down current of the
C=1 Chern insulator in b&c. Inset: zoom-in of curves at around IDC = 0 nA, which reveal the
threshold of EQAH break-down at ~ 1 nA. Note: Spike-like rapid changes in figure b-f are artifacts
from voltage source meter when switching output range.

You might also like