Download Fractional quantum Hall effect in suspended graphene probed with

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

Quantum computing wikipedia , lookup

Quantum electrodynamics wikipedia , lookup

Quantum teleportation wikipedia , lookup

Interpretations of quantum mechanics wikipedia , lookup

Orchestrated objective reduction wikipedia , lookup

Graphene wikipedia , lookup

Quantum machine learning wikipedia , lookup

Hydrogen atom wikipedia , lookup

Quantum key distribution wikipedia , lookup

Quantum group wikipedia , lookup

EPR paradox wikipedia , lookup

History of quantum field theory wikipedia , lookup

T-symmetry wikipedia , lookup

Hidden variable theory wikipedia , lookup

Quantum state wikipedia , lookup

Canonical quantization wikipedia , lookup

Transcript
Downloaded from http://rsta.royalsocietypublishing.org/ on May 4, 2017
Phil. Trans. R. Soc. A (2010) 368, 5403–5416
doi:10.1098/rsta.2010.0226
Fractional quantum Hall effect in suspended
graphene probed with two-terminal
measurements
BY I. SKACHKO1 , X. DU1,† , F. DUERR1 , A. LUICAN1 , D. A. ABANIN2, *,
L. S. LEVITOV3 AND E. Y. ANDREI1
1 Department
of Physics and Astronomy, Rutgers University, Piscataway,
NJ 08855, USA
2 Princeton Center for Theoretical Science, Princeton University, Princeton,
NJ 08544, USA
3 Department of Physics, Massachusetts Institute of Technology,
77 Massachusetts Avenue, Cambridge, MA 02139, USA
Recently, fractional quantization of two-terminal conductance was reported in suspended
graphene. The quantization, which was clearly visible in fields as low as 2 T and persistent
up to 20 K in 12 T, was attributed to the formation of an incompressible fractional
quantum Hall state. Here, we argue that the failure of earlier experiments to detect the
integer and fractional quantum Hall effect with a Hall-bar lead geometry is a consequence
of the invasive character of voltage probes in mesoscopic samples, which are easily shorted
out owing to the formation of hot spots near the edges of the sample. This conclusion is
supported by a detailed comparison with a solvable transport model. We also consider,
and rule out, an alternative interpretation of the quantization in terms of the formation
of a p–n–p junction, which could result from contact doping or density inhomogeneity.
Finally, we discuss the estimate of the quasi-particle gap of the quantum Hall state. The
gap value, obtained from the transport data using a conformal mapping technique, is
considerably larger than in GaAs-based two-dimensional electron systems, reflecting the
stronger Coulomb interactions in graphene.
Keywords: fractional quantum Hall effect; correlated electron systems; quantum transport
1. Introduction
Since its first isolation (Novoselov et al. 2004), graphene, a one-atom-thick
membrane of crystalline carbon, has broken many records with its unusual
physical properties: it is the stiffest material known, it has exquisite chemical
sensitivity, enabling detection of a single adsorbed atom, and, owing to
the relativistic nature of its charge carriers, the highest room temperature
*Author for correspondence ([email protected]).
† Present address: SUNY at Stony Brook, Department of Physics, Stony Brook, NY, USA.
One contribution of 12 to a Theme Issue ‘Electronic and photonic properties of graphene layers
and carbon nanoribbons’.
5403
This journal is © 2010 The Royal Society
Downloaded from http://rsta.royalsocietypublishing.org/ on May 4, 2017
5404
I. Skachko et al.
mobility known (Castro Neto et al. 2009). Yet, despite strong unscreened
Coulomb forces between these carriers, the expected manifestations of collective
behaviour (Apalkov & Chakraborty 2006; Peres et al. 2006; Toke et al. 2006;
Yang et al. 2006; Goerbig & Regnault 2007; Khveshchenko 2007; Wang et al.
2008) have been conspicuously difficult to detect. Here, we discuss requirements
for the sample preparation and measurement geometry that are necessary to
observe correlated electron behaviour such as the fractional quantum Hall effect
(FQHE). One condition is elimination of substrate-induced perturbations that
introduce fluctuations on energy scales larger than those characteristic for
electron correlations, which can be accomplished using suspended graphene (SG).
In addition, owing to the mesoscopic character of SG samples, the measurement
geometry and placing of probes are critical for the ability to observe the integer
quantum Hall effect (IQHE) and FQHE in SG (Du et al. 2008, 2009; Bolotin et al.
2009; Abanin et al. 2010).
The IQHE is a property of non-interacting two-dimensional electron systems
(2DESs) in a perpendicular magnetic field, B, marked by plateaus in the
transverse (Hall) conductance, sxy = ne 2 /h. Here, e is the electron charge, h is
Planck’s constant, n = ns h/eB is the filling factor and ns is the carrier density. For
graphene, the sequence of IQHE plateaus is anomalous, with n = ±4(n + 1/2),
n = 0, 1, 2, . . ., reflecting the fourfold spin and valley degeneracy, the electron–
hole symmetry and the number of filled Landau levels, n. The 1/2 shift is a
unique property of the relativistic charge carriers in graphene (Castro Neto et al.
2009). Correlations between electrons can lead to additional integer QH plateaus
outside this sequence corresponding to lifting of the spin or valley degeneracy.
For non-relativistic 2DESs, a strong correlation between the quasi-particles
gives rise to the FQHE (see Chakraborty & Pietilainen (1995) and references
therein) manifested as new plateaus at fractional filling: n = m/(2pm ± 1), where
p and m are integers. The FQHE reflects the formation of an incompressible
condensate that can be viewed as the IQHE for composite particles consisting
of an electron with 2p flux lines attached to it (Jain 1989). Excitations from
this condensate produce fractionally charged quasi-particles, with charge q ∗ =
e/(2pm ± 1), at an energy cost of the excitation gap, D. The gap can be estimated
as D = bEc , with Ec = e 2 /4p30 3lB the Coulomb energy, 30 the permittivity of free
space, 3 the effective dielectric constant, lB = (h̄/eB)1/2 the magnetic length and
b a numerical constant that depends on the filling factor. For n = 1/3, b ∼ 0.1
according to theory but is much smaller (approx. 0.01) in actual measurements
(Boebinger et al. 1987).
The FQHE in graphene is expected to deviate in important ways from that
in GaAs-based 2DESs. First, electrons in graphene are considerably more two
dimensional than in GaAs quantum wells, where the well widths range from
10 to 30 nm. This makes the interaction at short distances in graphene much
stronger than in quantum wells. The interactions in SG are further enhanced
owing to the absence of substrate screening (3 ∼ 1) in contrast to GaAs where
3 ∼ 13. Stronger interaction leads to a larger energy gap (Apalkov & Chakraborty
2006), and thus higher temperatures at which the FQHE can be observed.
Another difference arises from the fourfold spin and valley degeneracy. Because
of this degeneracy, the situation in graphene is more similar to that realized
in double quantum wells rather than in single quantum-well systems. However,
electron interactions in graphene correspond to nearly identical intra- and
Phil. Trans. R. Soc. A (2010)
Downloaded from http://rsta.royalsocietypublishing.org/ on May 4, 2017
Fractional quantum Hall effect
5405
inter-well interactions, a regime nearly impossible to achieve in GaAs systems.
This leads to an SU(4) symmetry of the interaction Hamiltonian, and the
prediction of new FQHE states with no analogue in GaAs (Toke et al. 2006;
Wang et al. 2008).
In the most common graphene devices, fabricated by mechanical exfoliation of
graphite onto an Si substrate capped with SiO2 , the IQHE is extremely robust
and can persist up to room temperature (Peres et al. 2006). In contrast, the
FQHE was never observed in these devices, in spite of intense efforts. Its absence
was attributed to substrate intrusiveness that, by introducing fluctuations on
energy scales larger than the correlations, can prevent the formation of the quite
delicate FQHE states. Based on the historical precedent in the GaAs-based 2DESs
where the observation of the FQHE became possible only after achieving better
material quality (see Chakraborty & Pietilainen (1995) and references therein),
it was expected that once the substrate disturbances were eliminated, correlation
effects would also be seen in graphene.
Recently, a significant improvement in sample quality was demonstrated in SG,
where all substrate-induced perturbations were eliminated (Bolotin et al. 2008;
Du et al. 2008). The combination of ballistic transport and low carrier density
that was achieved in these samples has raised hopes to observe the correlated
behaviour of the relativistic charge carriers. Surprisingly, measurements of the
Hall conductance in SG using the standard Hall-bar lead geometry failed to
exhibit the FQHE. In fact, not only was the FQHE not observed, but even the
normally much more robust IQHE was absent (Bolotin et al. 2008). By contrast,
as we show below, measurements of the SG samples using a two-terminal lead
geometry do exhibit both the IQHE and the FQHE.
2. Experimental system: failure of multi-terminal measurements
In the present study, the magneto-transport of SG devices was investigated in
both two-terminal and Hall-bar lead geometries. The SG devices were fabricated
from conventional devices, mechanically exfoliated onto Si/SiO2 substrates, by
removing the SiO2 layer with chemical etching (Du et al. 2008). In the final device,
the graphene sample is suspended from the Au/Ti leads (figure 1a and inset figure
2a). An estimate of the contact resistance of these leads was obtained through a
multi-terminal measurement of non-suspended samples with leads having contact
area and geometry similar to that in the SG samples. Its value, typically
approximately 200 U, is significantly less than the quantum resistance with no
measurable gate-voltage dependence. Following fabrication, the samples were
baked at 200◦ C in flowing forming gas (Ar : H2 , 9 : 1) for 2 h and current annealed
(Moser et al. 2007) in vacuum at low temperature to further improve sample
quality. To ensure mechanical integrity, we focused on short devices, less than
1 mm long. Longer devices tend to collapse during fabrication or when subject to a
gate voltage. The carrier density in the suspended devices, ns ∼ 2 × 1014 Vg (m−2 ),
was controlled by the gate voltage Vg (measured in volts; Du et al. 2008).
In figure 1, we show the results obtained for an SG device with a Hall-bar lead
geometry. The filling factor dependence of the Hall and longitudinal conductance
is shown in figure 1b. Surprisingly, Gxy does not show clear plateaus, and its
value is significantly lower than that expected for the QHE plateaus, ne 2 /h. At
Phil. Trans. R. Soc. A (2010)
Downloaded from http://rsta.royalsocietypublishing.org/ on May 4, 2017
5406
I. Skachko et al.
(b) 8
(a)
Gxx, Gxy (e2/h)
graphene
6
4
Gxx
2
0
T = 1.2 K
B=1T
Gxy
–2
–4
4
(d)
(c)
Vs = V
0
n
Vd = 0
Vs = V
Vd = 0
I
Figure 1. (a) Scanning electron microscope (SEM) image of SG sample with Hall-bar lead geometry.
(b) Longitudinal and Hall conductance for the sample in (a) obtained from the measured resistances
2 + R 2 ), G = R /(R 2 + R 2 ). (c) Equipotential line distribution in a sample
Gxx = Rxx /(Rxx
xy
xy
xy
xx
xy
with W /L = 1.78, for a large Hall angle (sxy /sxx = 20) illustrating the regions of sharp potential
drop (hot spots) at opposite corners of the sample. Vs and Vd denote source and drain voltage.
(d) Shorting of Hall voltage by invasive leads. In the QH regime, the carriers move in ballistic edge
channels originating from one current lead and traversing a region of potential gradient, ‘hot spot’
(grey circle), before entering the opposite lead. A voltage lead covering the hot-spot region (hashed
rectangles) shorts out the Hall voltage as observed in (b).
the same time, Gxx , instead of vanishing for n corresponding to the QHE, develops
plateau-like features. This indicates that, although the QH effect does occur in
our system, the Hall voltage probes are short-circuited through the source and
drain leads.
Since in our devices, the size of the voltage probes is comparable to the sample
size, it is important to take into account the effects of geometry. The behaviour
of magneto-transport in such devices is to be contrasted with that in Hall-bar
devices, conventionally used for studying Hall transport. In long Hall bars, the
electric field is mostly transverse, and, as a result, the Hall voltage is practically
insensitive to the positions of the Hall probes. This can be understood in an
intuitive manner by considering the edge-state model of QH transport, in which
transport is due to one-dimensional chiral edge channels. This model predicts
that the voltage probes on the opposite sides of the sample measure the chemical
potential of the top (bottom) chiral edge state, which is equal to the source
(drain) potential. As a result, the measured Hall voltage is quantized under the
QH effect conditions.
Phil. Trans. R. Soc. A (2010)
Downloaded from http://rsta.royalsocietypublishing.org/ on May 4, 2017
5407
Fractional quantum Hall effect
(a)
(b)
106
5
R (h/e2)
T = 1.2 K
m (cm2 Vs–1)
r (kW)
15
10
1T
2T
3T
4T
5T
6T
8T
10 T
12 T
n = 1/3
1 n=1
n=2
n=6
n = 10
n = 14
105
1 10 100
ns (×109) cm–2
0.1
0
200
0
n (×109) cm–2
–200
–20
(c)
0
Vg (V)
20
(d)
105
n = 1/3
104
n=1
n=2
n=3
n=6
1T
2T
3T
4T
5T
6T
8T
10 T
12 T
10
6
2
–14
–10
R (W)
G (e2/h)
14
Vg = 0.92 V
103
n
–6
–2
0
1.08
2
3
4
0
2
4
B (T)
6
Vg = 5 V
10
15
17.5
20
8
Figure 2. (a) Comparison of measured density dependence of resistivity in zero field (grey) with
the calculated curve (black) for an ideal ballistic graphene junction with the same dimensions
(Tworzydlo et al. 2006). Upper inset: false-colour SEM image of SG device with L = 0.6 mm and
W = 1.4 mm. Lower inset: Drude mobility versus carrier density. (b) Gate-voltage dependence of
two-terminal resistance at T = 1.2 K for the indicated field values. Dashed lines mark the resistance
values on the QHE plateaus. In addition to the usual non-interacting IQHE, we note a plateau
at n = −1 and a newly observed one at n = −1/3, clearly visible on the hole side (Vg < 0). (c)
G(n) obtained from scaled electron density, ns h/eB, in (b). All the curves collapse together, giving
QH plateaus at G(n) = ne 2 /h centred at the correct filling factors in the hole sector. The N-shape
distortions characteristic of the two-terminal measurements with L < W are clearly seen (Abanin &
Levitov 2008). (d) Evolution of QHE resistance plateaus with magnetic field for indicated values of
gate voltage for an SG sample with W = 2.6 mm, L = 0.5 mm. Horizontal lines indicate theoretical
resistance values for QH plateaus.
In contrast, for the case of a sample with an aspect ratio of order one,
the existence of so-called hot spots situated on the sides of the sample
(Wakabayashi & Kawaji 1978; Klass et al. 1991; Komiyama et al. 2006; Ikushima
et al. 2007) near the contacts, where the component of the field parallel to the edge
is large, makes the measured Hall voltage strongly dependent on the arrangement
of the Hall probes. The voltage probe placed in the vicinity of the hot spot senses a
potential which is intermediate between that of source and drain. This can cause
a significant reduction in the measured Hall voltage between opposite voltage
probes. Thus, in contrast to the case of a long Hall bar, the measured Hall voltage
is not quantized.
Phil. Trans. R. Soc. A (2010)
Downloaded from http://rsta.royalsocietypublishing.org/ on May 4, 2017
5408
I. Skachko et al.
As a concrete model illustrating this behaviour, we consider the rectangular
geometry shown in figure 1c,d. In this case, the potential distribution can be
found exactly by the conformal-mapping technique (Rendell & Girvin 1981). In
the limit of large Hall angle, and for the sample aspect ratio of order one, the
distribution of potential and current becomes highly non-uniform, in contrast
to the more familiar case of a long Hall bar. This distribution is characterized
by two accumulation regions at the opposite corners of the samples (hot spots),
where field lines are concentrated. In the hot-spot corners, which are located
at the opposite corners (x = 0, y = 0), (x = L, y = W ), where the x-axis is
horizontal and y-axis is vertical, the electric field exhibits a power-law singularity
(Rendell & Girvin 1981),
pDZ −2q/p
iq
,
(2.1)
E ∝ e tan
2L
where q is the Hall angle, E = Ex − iEy is the complex electric field and DZ =
Dx + iDy is the complex distance to the corner of the sample. In the other two
corners, the field does not exhibit a divergence,
pDZ 2q/p
iq
.
(2.2)
E ∝ e tan
2L
The difference in the longitudinal electric field on the opposite sides of the sample,
evident from equations (2.1) and (2.2), will result in a large distortion of the
measured Hall voltage.
Another manifestation of current and field accumulation in the hot-spot
regions, which can affect the measured Hall voltage, is the strong heating of these
regions. Indeed, the Joule heating is strongly enhanced by the singular behaviour
of the electric field in the hot-spot equation (2.1) in the limit of the large Hall
angle. The heating can affect the microscopic conductivity tensor in the vicinity
of the hot spot, giving rise to the increased sxx in that region. The increased
conductivity may lead to the short-circuiting between the source (drain) and the
voltage probe places in the vicinity of the hot spot.
Another interesting mechanism, which may further increase conductivity near
the corners of the sample, can be due to a strain-induced pseudo-magnetic field
(Prada et al. 2010). The strain of the SG sheet, which may result, e.g. from the
electrostatic interaction with the gate (Prada et al. 2010), is strongest near the
corners; the pseudo-magnetic field is expected to exhibit a power-law singularity
near the corners. The estimated magnitude of the pseudo-magnetic field near the
corners, of the order of a few tesla, is comparable to the external magnetic field in
our experiments. The large pseudo-magnetic field makes the effective filling factor
non-uniform across the sample. Because of that, even when the filling factor in the
bulk of the sample corresponds to an incompressible QH state with vanishing sxx ,
the regions near the corners may be conducting, which would lead to the shortcircuiting between the source (drain) and voltage probes. This would decrease
the measured Hall voltage, which would no longer be quantized.
We believe that the absence of quantization in our six-terminal measurements is
owing to the effects associated with hot spots, discussed above. In our mesoscopic
samples, the lead dimensions are comparable to sample size. Because of that,
Phil. Trans. R. Soc. A (2010)
Downloaded from http://rsta.royalsocietypublishing.org/ on May 4, 2017
Fractional quantum Hall effect
5409
it is likely that the shorting of the Hall voltage observed in SG devices is because
of the fact that the voltage leads were placed within the hot-spot regions. In
future experiments, this difficulty may be circumvented by fabricating samples
with a larger aspect ratio L/W > 1 and placing voltage probes further away from
the hot-spot regions.
3. Two-terminal measurements and fractional quantum Hall effect
Although, at present, reliable Hall-bar measurements are not feasible in SG, it
is still possible, as we show in the next section, to obtain the Hall conductance
on QH plateaus as well as the longitudinal conductivity, sxx , from a two-terminal
conductance (G) measurement, where the leads completely avoid the hot spots.
The results of the two-terminal measurements on SG samples are shown
in figure 2. From the zero-field resistivity (figure 2a), we extract the Drude
mobility, m = (rne)−1 , and find that its density dependence, m ∝ ns−1/2 , closely
follows that of a ballistic device with the same dimensions (figure 2a, lower
inset) for hole carriers (negative Vg ), with a somewhat less good agreement
on the electron side. In a finite magnetic field, the two-terminal conductance
exhibits well-defined plateaus, G = |n|e 2 /h for n = ±2, ±6, ±10, ±14, . . . , seen
already below 1 T (figure 2b). In between the plateaus, we observe conductance
maxima consistent with theory for our device geometry (W /L > 1) (Abanin &
Levitov 2008; Williams et al. 2009). When G is plotted versus the filling factor,
the curves for all fields lie on top of each other (figure 2c) and exhibit a series of
well-defined IQHE plateaus, consistent with the high carrier mobility (Abanin &
Levitov 2008) and negligible contact resistance. Above 2 T, additional plateaus
develop at n = ±1, 3, reflecting interaction-induced lifting of the degeneracy
in the n = 0 and n = 1 Landau levels. This is in sharp contrast to earlier
measurements on non-suspended samples, where the n = ±1 plateaus were only
seen above 26 T (Zhang et al. 2006) and the one at n = 3 was absent. At low
temperatures, we observe an FQHE plateau at n = −1/3 with G = (1/3)e 2 /h,
which becomes better defined with increasing field (figure 2b). Zooming into
the low n regime in figure 3a, we note that the plateaus at n = −1/3, −1 show
accurate values of the quantum Hall conductance. Crucially, the densities at which
these plateaus occur scale linearly with the B field, ns = neB/h, corresponding
to constant filling factor values. The FQHE at n = −1/3 is quite robust; it
appears already at approximately 2 T at low temperatures and persists up
to approximately 20 K at 12 T. At even lower filling factors, the conductance
becomes vanishingly small, indicating that the FQHE in graphene competes with
an insulating phase at n = 0. The properties of this phase including temperature
dependence and activation gaps are discussed in Du et al. (2009). It is not obvious
a priori that the correlated state leading to the FQHE for the relativistic charge
carriers in graphene is the same as that for the 2DESs in semiconductors. In
order to gain insight into the nature of the FQHE state in graphene, it is
desirable to obtain an accurate measure of the excitation gap. Experimentally,
the gap is usually extracted from the temperature dependence of sxx that
can be obtained, together with sxy , from a multi-probe measurement, such as
a Hall-bar lead geometry. However, as discussed above, this method fails in
mesoscopic samples.
Phil. Trans. R. Soc. A (2010)
Downloaded from http://rsta.royalsocietypublishing.org/ on May 4, 2017
5410
I. Skachko et al.
G (e2/h)
(a)
(b)
6T
8T
10 T
12 T
40 K
20 K
10 K
4K
1.2 K
1
1/3
T = 1.2 K
–1
n
–1/3
0
–1
n
–1/3
0
Figure 3. Field and temperature dependence of the n = −1/3, −1 plateaus. (a) Zoom into the low n
segment of the data in figure 2c. Well-defined conductance plateaus at G = e 2 /h and G = 1/3(e 2 /h)
occurring at the precise filling factors n = −1, −1/3 are already visible at 2 T. (b) Temperature
dependence of the QH plateaus. On the plateaus, the conductance increases with temperature
faster than off the plateaus. Both plateau features remain visible up to T ∼ 20 K.
4. Fractional quantum Hall effect versus integer quantum Hall effect
Here, we discuss to what extent the interpretation of the fractional quantization
in terms of a Laughlin FQH state is unambiguous. Recently, it was found that
integer QHE in locally gated graphene devices, such as p–n and p–n–p junctions,
can manifest itself in fractional quantization of the two-terminal conductance
(Ozyilmaz et al. 2007; Williams et al. 2007). Such fractional quantization was
observed for the spatially varying density profile in which incompressible QH
states with different integer filling factors n, n∗ are created in different parts of
the device. The fractional quantization was understood (Abanin & Levitov 2007)
in terms of edge-state transport, in which the QH edge states propagate along
sample edges as well as along p–n, p–p and n–n boundaries.
Thus, a natural question arises whether our observations can be explained in
terms of the formation in SG of a p–n–p junction system. In this section, we shall
investigate whether a model involving a p–n–p system with integer QH states
that result from the lifted degeneracies of Landau levels (LLs) could account for
the fractional conductance observed in our experiment. On very general grounds,
we conclude that such a model is inconsistent with our observations, which rules
out the p–n–p junction scenario, and suggests that, in fact, a genuine fractional
QH effect has been observed.
Before we proceed, let us briefly discuss the possible origin of a p–n–p junction
in our graphene devices. The most likely mechanism of p–n–p junction formation
is contact doping (Giovannetti et al. 2008; Huard et al. 2008). The proximity
to metal is expected to pin the electron density below the contact to either a
positive or negative value, depending on the type of the metal (Giovannetti et al.
2008). Therefore, generally the density across the device is non-uniform, with
p–n, n–n or p–p junctions formed in the vicinity of the contacts. The resulting
electron–hole asymmetry of the transport has been verified experimentally
(Huard et al. 2008). Alternatively, p–n–p junctions in our experiments may stem
from density inhomogeneity induced by disorder potential.
Phil. Trans. R. Soc. A (2010)
Downloaded from http://rsta.royalsocietypublishing.org/ on May 4, 2017
Fractional quantum Hall effect
5411
The conductance of a p–n–p junction, analysed in Ozyilmaz et al. (2007) for
incompressible QH states, was shown to be a simple function of filling factors in
the inner and outer regions,
G=
e2
|nn∗ |
,
|n∗ | + |n∗ − n| h
(4.1)
where n is the filling factor in the inner region, and n∗ is the filling factor
in the outer region. A similar result can be obtained in the ideal Hall effect
limit (large Hall angle; using a method similar to that used in Abanin &
Levitov (2008)). In the regime where LL degeneracies are lifted, such that
n, n∗ = 0, ±1, ±2, ±3, . . ., this formula predicts the following quantized values:
G = 1/3, 2/3, 2/5, 1/2, . . . , e 2 /h.
The quantized value of G = (1/3)(e 2 /h), which is most interesting to us, occurs
when n = 1, n∗ = −1 or n = −1, n∗ = 1. We consider the situation where the p–n–p
junction is formed by contact doping, and model the behaviour or n, n∗ as a
function of back-gate voltage Vbg as follows: n = aVbg and n∗ = aVbg + bVc , where
Vc is a parameter that describes the contact doping. In that, assuming that Vc
is constant and Vbg is varying, n and n∗ change sign at different values of Vbg . As
a result, in the region where n, n∗ have opposite signs, the behaviour of G(Vbg ),
predicted by equation (2.1) is non-monotonic. Hence, the one-third plateau, which
is realized when n and n∗ have different signs, is positioned between two plateaus
with zero conductance, G = 0, which occur when n = 0 or n∗ = 0. Similar behaviour
will arise when Vc depends smoothly on Vbg . Such non-monotonic behaviour is
inconsistent with our experiment, where only one G = 0 plateau is observed, and
function G(Vbg ) is monotonic.
To further illustrate this point, we model the behaviour of the conductance of
a p–n–p junction in the LL-split regime as a function of parameters Vbg , Vc . We
have employed a method similar to that used in Abanin & Levitov (2008). The
result is illustrated in figure 4, where the experimentally measured conductance
trace is given by the vertical cross section (dashed line) of a two-dimensional plot,
with a non-monotonic behaviour around one-third. It is clear from the figure that
allowing Vc to vary with Vbg would not change these conclusions.
Another discrepancy with the p–n–p model, which is evident from figure 4,
is that the fractional conductance plateaus are positioned at Vbg values of Vbg
that correspond to integer filling factors in the central region. By contrast, in our
experiment, the conductance behaviour is monotonic, with fractional plateaus
found at values of Vbg corresponding to fractional filling factors. Thus, we
conclude that the p–n–p junction formation is inconsistent with observations.
The only viable explanation of the observed fractional conductance, in our view,
is that a spatially uniform FQH state is realized in our SG samples.
5. Extracting the longitudinal and transverse conductivity from
two-terminal conductance
Since, at present, there are no reliable Hall-bar measurements in SG, here, we
attempt to obtain sxx and sxy from the two-terminal conductance G. Since this
quantity depends on sxx , sxy and sample geometry, ‘deconvolving’ G requires
an additional input. Such input is provided by the conformal invariance of the
Phil. Trans. R. Soc. A (2010)
Downloaded from http://rsta.royalsocietypublishing.org/ on May 4, 2017
5412
I. Skachko et al.
2.0
2/3
2/5
Vb (arbitrary units)
2
1
1/3
1.5
1
1/2
0
1
2
1.0
1/3
1
1/2
0.5
2/3
Vc (arbitrary units)
2/5
0
Figure 4. Conductance of a p–n–p junction system versus the back-gate voltage Vbg , and parameter
Vc that describes doping by contacts, modelled using a method similar to that of Abanin &
Levitov (2008). Splitting of the fourfold LL degeneracy by the interactions is accounted for by
integer incompressible filling factor values n, n∗ = 0, ±1, ±2, ±3, . . . . This gives fractional quantized
conductance values, including G = 1/3, 2/3, 2/5, 1/2, . . . , e 2 /h (labelled by numbers in boxes). The
measured conductance trace is taken along a cross section of this plot Vc = Vc (Vbg ). Vertical cross
section Vc = constant is indicated by a grey dashed line. The behaviour of the conductance is
non-monotonic around the one-third plateau, which is in clear departure from observations.
magneto-transport problem (Abanin & Levitov 2008). In this approach, sxx and
sxy are interpreted as a real and imaginary part of a complex number s =
sxx + isxy , and thereupon the transport equations become conformally invariant.
Applied to a rectangular two-lead geometry, such as that in our experiment,
theory (Abanin & Levitov 2008) yields a specific dependence of G on sxx , sxy and
the aspect ratio W /L. Interestingly, the same dependence describes two-terminal
conductance for an arbitrary sample shape, with the ‘effective aspect ratio’ W /L
encoding the geometry dependence.
There are several ways to use this approach for determining sxx . One is to
focus on the plateaus where sxx is small and sxy is quantized, sxy = ne 2 /h.
Expanding G in the small ratio sxx /|sxy | 1, the deviation from a quantized value
can be expressed as G = |n|e 2 /h + ksxx . The coefficient k, obtained from theory
(Abanin & Levitov 2008), as a function of W /L, is positive (negative) for L < W
(L > W ) and zero for L = W . Despite its conceptual simplicity, we found this
approach not quite adequate, since the effective W /L may significantly deviate
from the geometric aspect ratio of the sample (Williams et al. 2009). Further, since
sxx and sxy change with n, the conductance plateaus exhibit N-shaped distortions,
rendering the deviation in G from a quantized value unsuitable for an accurate
estimate of sxx .
Considerably more reliable results can be obtained by focusing on the N-shaped
distortions of QH plateaus (Abanin & Levitov 2008; Williams et al. 2009). The
N-shaped features can be described by the density-dependent sxx (n) and sxy (n)
obtained from the so-called semicircle relation (see Ruzin & Feng (1995) and
references therein). This relation, which is also rooted in the complex-variable
Phil. Trans. R. Soc. A (2010)
Downloaded from http://rsta.royalsocietypublishing.org/ on May 4, 2017
5413
Fractional quantum Hall effect
(a)
G, sxx(e2/h)
2
T = 1.2 K
B = 12 T
1
(b)
sxx (n = 1)
fit
sxx (n = 1/3)
fit
0.3
sxx(e2/h)
G data
G theory
sxx calculated
0.2
B = 12 T
0.1
1/3
–1
n
0
0.2
0.4
1/T
0.6
0.8
Figure 5. (a) Theoretical fit of two-terminal conductance (dark grey) compared with data (black)
using the semicircle model described in the text. The longitudinal conductivity sxx (light grey),
obtained from the best fit, has peaks at the plateau-to-plateau transitions and minima on the
plateaus. Values at the minima are used to estimate sxx for incompressible QH states. (b) Fits
of the temperature dependence of sxx at n = −1/3, −1 with a thermal activation model, sxx ∼
exp(−D/2kB T ). We find Dn=1/3 ∼ 4.4 K and Dn=1/3 ∼ 9.6 K.
interpretation of magneto-transport, is known to work well in GaAs-based 2DESs,
both in the integer and in the fractional QHE regimes (Shahar et al. 1997; Murzin
et al. 2002, 2005). For a plateau-to-plateau transition between incompressible
2
= (sxy − n1 )(n2 − sxy ) (in
filling factors n1 < n2 , the semicircle relation gives sxx
2
units of e /h). For fitting the conductance data shown in figure 3 with the
incompressible states at n = 0, −1/3, −1, −2, we model the contribution of each
2
QH transition by a Gaussian, sxx (n) = (1/2)(n2 − n1 )e−A(n−nc ) , n1 < nc < n2 , and
find the corresponding sxy from the semicircle relation. This gives a contribution
to the longitudinal and Hall conductivity of each of the relevant Landau (sub)
levels. The net conductivity, found as a sum of such independent contributions,
is then used to calculate G(n) using the approach (Abanin & Levitov 2008)
discussed above.
We treat the sxx peak positions and widths, as well as the effective ratio W /L,
as variational parameters. As illustrated in figure 5a, the Gaussian model with
individually varying peak widths and positions is found to provide a rather good
description of the data, at the same time giving the quantity sxx (n). The approach
based on treating W /L as a variational parameter, in general different from the
actual sample aspect ratio, works rather well in the integer QH regime (Williams
et al. 2009). It was conjectured (Williams et al. 2009) that variations in the bestfit value of W /L account for the sample-dependent specifics of the current flow
pattern, such as those owing to imperfect contacts and/or contact doping.
In order to extract sxx at the plateaus n = −1/3 and n = −1, we analysed a set
of fits that best follow the data near these filling factors (figure 5a). In both cases,
we found an optimal effective aspect ratio W /L ≈ 1.7 (such deviation from the
geometric aspect ratio, which for our sample is close to 2.5, is consistent with the
results of Williams et al. (2009)). Best fits were found from the standard deviation
in the conductance averaged over the range of densities on the plateau and around
it; in the case of the one-third plateau, we excluded a small interval near n ≈ 0.4,
which we believe is related to another incipient fractional QHE feature.
Phil. Trans. R. Soc. A (2010)
Downloaded from http://rsta.royalsocietypublishing.org/ on May 4, 2017
5414
I. Skachko et al.
The value of sxx at the minima of sxx (n) was taken as an estimate of the
longitudinal conductivity of the incompressible QH states. Statistical error was
estimated from the spread in the sxx values found from the fits with standard
deviation within 30 per cent of the best fit. The resulting error bar, displayed
in figure 5b, is below 10 per cent at 10 K, and increases to 20–25% at 1.2 K as a
result of pronounced mesoscopic fluctuations developing at low temperatures.
We used the 12 T data at T = 1.2, 4.2, 6 and 10 K to infer the
temperature dependence of sxx in the n = −1/3 and n = −1 states. The results,
displayed in figure 5b, were analysed with a thermal activation model, sxx ∼
exp(−D/2kB T ). For n = −1/3, we find Dn=1/3 ∼ 4.4 K. We also considered fits
to the variable-range-hopping dependence, sxx ∼ exp(−(T ∗ /T )0.5 ), but have not
found a discernible statistical advantage over the activation dependence. This
gap is close to 2.5 times larger than the gap measured in the best samples in
the GaAs/GaAs-based 2DESs. Theoretical estimates (Toke et al. 2006) for the
gap in the n = 1/3 state in SG gives Dn=1/3 ∼ 0.1Ec . Taking the random-phase
approximation result for the intrinsic screening function of graphene in zero field
(Gonzalez et al. 1999), 3 ∼ 5.24, the theoretical value of the gap at B = 12 T is
Dn=1/3 ∼ 42 K. This implies that it is possible to observe an even more robust
FQHE at n = −1/3 and other fractions by improving the quality of the graphene
samples. By contrast, in the GaAs/GaAs-based 2DESs, the gaps are one order
of magnitude smaller than in graphene owing to the higher dielectric screening
and the finite width of the quantum wells. In the highest quality semiconductor
samples, the theoretical gap values are already achieved at high fields so that not
much improvement in the robustness of the FQHE can be expected there.
For the n = −1 state, we find that activated behaviour, sxx ∼ exp(−D/2kB T ),
with the energy gap value D/kB = 9.6 K, best fits the data. This is only 5 per cent
of the theoretically predicted value (Toke et al. 2006), and we expect it would
increase in higher quality graphene samples.
6. Summary
In this work, we analyse the magneto-transport measurements in SG, in which
electrons are isolated from environmental disturbances caused by the substrate.
We demonstrate that by using two-terminal measurement geometry, the effect
of invasive voltage probes can be eliminated, making the collective behaviour of
the relativistic two-dimensional carriers readily observable. We present a detailed
analysis of recent magneto-transport measurements and show that SG samples
exhibit a robust FQHE at n = 1/3.
From the temperature dependence of the longitudinal conductivity, we found
the energy gap of the quasi-particle excitations. Its magnitude, while only
10 per cent of the theoretically predicted value indicating imperfect sample
quality, is still more than three times larger than that for the best semiconductorbased 2DESs. The robust character of the FQHE observed in graphene reflects
enhancement of the Coulomb interactions owing to lower dielectric screening and
to the more two-dimensional nature of electrons in suspended graphene.
E.Y.A. acknowledges DOE support under DE-FG02-99ER45742 and partial NSF support under
NSF-DMR-0906711. L.S.L. acknowledges ONR support under N00014-09-1-0724.
Phil. Trans. R. Soc. A (2010)
Downloaded from http://rsta.royalsocietypublishing.org/ on May 4, 2017
Fractional quantum Hall effect
5415
References
Abanin, D. A. & Levitov, L. S. 2007 Quantized transport in graphene p-n junctions in a magnetic
field. Science 317, 641–643. (doi:10.1126/science.1144672)
Abanin, D. A. & Levitov, L. S. 2008 Conformal invariance and shape-dependent conductance of
graphene samples. Phys. Rev. B 78, 035416. (doi:10.1103/PhysRevB.78.035416)
Abanin, D. A., Skachko, I., Du, X., Andrei, E. Y. & Levitov, L. S. 2010 Fractional quantum Hall
effect in suspended graphene: transport coefficients and electron interaction strength. Phys. Rev.
B. 81, 115410. (doi:10.1103/PhysRevB.81.115410)
Apalkov, V. M. & Chakraborty, T. 2006 Fractional quantum Hall states of Dirac electrons in
graphene. Phys. Rev. Lett. 97, 126801. (doi:10.1103/PhysRevLett.97.126801)
Boebinger, G. S., Stormer, H. L., Tsui, D. C., Chang, A. M., Hwang, J. C. M., Cho, A. Y., Tu,
C. W. & Weimann, G. 1987 Activation energies and localization in the fractional quantum Hall
effect. Phys. Rev. B 36, 7919–7929. (doi:10.1103/PhysRevB.36.7919)
Bolotin, K. I., Sikes, K. J., Jiang, Z., Fudenberg, G., Hone, J., Kim, P. & Stormer, H. L.
2008 Ultrahigh electron mobility in suspended graphene. Solid State Commun. 146, 351–355.
(doi:10.1016/j.ssc.2008.02.024)
Bolotin, K. I., Ghahari, F., Shulman, M. D., Stormer, H. L. & Kim, P. 2009 Observation of the
fractional quantum Hall effect in graphene. Nature 462, 196–199. (doi:10.1038/nature08582)
Castro Neto, A. H., Guinea, F., Peres, N. M. R., Novoselov, K. S. & Geim, A. K. 2009 The electronic
properties of graphene. Rev. Mod. Phys. 81, 109. (doi:10.1103/RevModPhys.81.109)
Chakraborty T. & Pietilainen, P. (eds) 1995 The quantum Hall effects: integral and fractional,
vol. 85. Series in Solid State Physics. New York, NY: Springer.
Du, X., Skachko, I., Barker, A. & Andrei, E. Y. 2008 Approaching ballistic transport in suspended
graphene. Nat. Nanotechnol. 3, 491–495. (doi:10.1038/nnano.2008.199)
Du, X., Skachko, I., Duerr, F., Luican, A. & E. Y Andrei 2009 Fractional quantum Hall effect and
insulating phase of Dirac electrons in graphene. Nature 462, 192–195. (doi:10.1038/nature08522)
Giovannetti, G., Khomyakov, P. A., Brocks, G., Karpan, V. M., van den Brink, J. & Kelly, P.
J. 2008 Doping graphene with metal contacts. Phys. Rev. Lett. 101, 026803. (doi:10.1103/Phys
RevLett.101.026803)
Goerbig, M. O. & Regnault, N. 2007 Analysis of a SU(4) generalization of Halperin’s wave function
as an approach towards a SU(4) fractional quantum Hall effect in graphene sheets. Phys. Rev.
B 75, 241405(R). (doi:10.1103/PhysRevB.75.241405)
Gonzalez, J., Guinea, F. & Vozmediano, M. A. H. 1999 Marginal-fermi-liquid behavior
from two-dimensional Coulomb interaction. Phys. Rev. B 59, R2474–R2477. (doi:10.1103/
PhysRevB.59.R2474)
Huard, B., Stander, N., Sulpizio, J. A. & Goldhaber-Gordon, D. 2008 Evidence of the role of
contacts on the observed electron-hole asymmetry in graphene. Phys. Rev. B 78, 121402(R).
(doi:10.1103/PhysRevB.78.121402)
Ikushima, K., Sakuma, H., Komiyama, S. & Hirakawa, K. 2007 Visualization of quantum Hall
edge channels through imaging of terahertz emission. Phys. Rev. B. 76, 165323. (doi:10.1103/
PhysRevB.76.165323)
Jain, J. K. 1989 Composite-fermion approach for the fractional quantum Hall effect. Phys. Rev.
Lett. 63, 199–202. (doi:10.1103/PhysRevLett.63.199)
Khveshchenko, D. V. 2007 Composite Dirac fermions in graphene. Phys. Rev. B 75, 153405.
(doi:10.1103/PhysRevB.75.153405)
Klass, U., Dietsche, W. & Klitzing, K. V. 1991 Image of the dissipation in gated quantum Hall
effect samples. Z. Phys. B 82, 351–354. (doi:10.1007/BF01357178)
Komiyama, S., Sakuma, H., Ikushina, K. & Hirakawa, K. 2006 Electron temperature of hot spots
in quantum Hall conductors. Phys. Rev. B 73, 045333. (doi:10.1103/PhysRevB.73.045333)
Moser, J., Barreiro, A. & Bachtold, A. 2007 Current-induced cleaning of graphene. Appl. Phys.
Lett. 91, 163513. (doi:10.1063/1.2789673)
Murzin, S. S., Weiss, M., Jansen, A. G. M. & Eberl, K. 2002 Universal flow diagram for the
magnetoconductance in disordered GaAs layers. Phys. Rev B 66, 233314. (doi:10.1103/Phys
RevB.66.233314)
Phil. Trans. R. Soc. A (2010)
Downloaded from http://rsta.royalsocietypublishing.org/ on May 4, 2017
5416
I. Skachko et al.
Murzin, S. S., Dorozhkin, S. I., Maude, D. K. & Jansen, A. G. 2005 Scaling flow diagram in the
fractional quantum Hall regime of GaAs/AlxGa1-xAs heterostructures. Phys. Rev. B 72, 195317.
(doi:10.1103/PhysRevB.72.195317)
Novoselov, K. S., Geim, A. K., Morozov, S. V., Jiang, D., Zhang, Y., Dubonos, S. V., Grigorieva,
I. V. & Firsov, A. A. 2004 Electric field effect in atomically thin carbon films. Science 306,
666–669. (doi:10.1126/science.1102896)
Ozyilmaz, B., Jarillo-Herrero, P., Efetov, D., Abanin, D. A., Levitov, L. S. & Kim, P. 2007
Electronic transport and quantum Hall effect in bipolar graphene p-n-p junctions. Phys. Rev.
Lett. 99, 166804. (doi:10.1103/PhysRevLett.99.166804)
Peres, N. M. R., Guinea, F. & Castro Neto, A. H. 2006 Electronic properties of disordered twodimensional carbon. Phys. Rev. B 73, 125411. (doi:10.1103/PhysRevB.73.125411)
Prada, E., San-Jose, P., León, G., Fogler, M. M. & Guinea, F. 2010 Singular elastic strains
and magnetoconductance of suspended graphene. Phys. Rev. B 81, 161402(R). (doi:10.1103/
PhysRevB.81.161402)
Rendell, R. W. & Girvin, S. M. 1981 Hall voltage dependence on inversion-layer geometry in the
quantum Hall-effect regime. Phys. Rev. B 23, 6610–6614. (doi:10.1103/PhysRevB.23.6610)
Ruzin, I. & Feng, S. 1995 Universal relation between longitudinal and transverse conductivities in
quantum Hall effect. Phys. Rev. Lett. 74, 154–157. (doi:10.1103/PhysRevLett.74.154)
Shahar, D., Tsui, D. C., Shayegan. M., Shimshoni, E. & Sondhi, S. L. 1997 A different view of
the quantum Hall plateau-to-plateau transitions. Phys. Rev. Lett. 79, 479–482. (doi:10.1103/
PhysRevLett.79.479)
Toke, C., Lammert, P. E., Jain, J. K. & Crespi, V. H. 2006 Fractional quantum Hall effect in
graphene. Phys. Rev. B 74, 235417. (doi:10.1103/PhysRevB.74.235417)
Tworzydlo, J., Trauzettel, B., Titov, M., Rycerz, A. & Beenakker, C. W. J. 2006 Sub-Poissonian
shot noise in graphene. Phys. Rev. Lett. 96, 246802. (doi:10.1103/PhysRevLett.96.246802)
Wakabayashi, J. & Kawaji, S. 1978 Hall effect in Silicon MOS inversion layers under strong
magnetic fields. J. Phys. Soc. Jpn. 44, 1839–1849. (doi:10.1143/JPSJ.44.1839)
Wang, H., Sheng, D. N., Sheng, L. & Haldane, F. D. M. 2008 Broken-symmetry states of Dirac
fermions in graphene with a partially filled high landau level. Phys. Rev. Lett. 100, 116802.
(doi:10.1103/PhysRevLett.100.116802)
Williams, J. R., DiCarlo, L. & Marcus, C. M. 2007 Quantum Hall effect in a gate-controlled p-n
junction of graphene. Science 317, 638–641. (doi:10.1126/science.1144657)
Williams, J. R., Abanin, D. A., DiCarlo, L., Levitov, L. S. & Marcus, C. M. 2009 Quantum
Hall conductance of two-terminal graphene devices. Phys. Rev. B 80, 045408. (doi:10.1103/
PhysRevB.80.045408)
Yang, K., Das Sarma, S. & MacDonald, A. H. 2006 Collective modes and skyrmion excitations
in graphene SU(4) quantum Hall ferromagnets. Phys Rev B. 74, 075423. (doi:10.1103/
PhysRevB.74.075423)
Zhang, Y. et al. 2006 Landau-level splitting in graphene in high magnetic fields. Phys. Rev. Lett.
96, 136806. (doi:10.1103/PhysRevLett.96.136806)
Phil. Trans. R. Soc. A (2010)